Skip to main content

Full text of "Evolution of Neutron-Star, Carbon-Oxygen White-Dwarf Binaries"

See other formats


Evolution of Neutron-Star, Carbon-Oxygen White-Dwarf Binaries 



G. E. Brown, C.-H. Lee 
Department of Physics & Astronomy, State University of New York, Stony Brook, New 

York 11794, USA 
Korea Institute for Advanced Study, Seoul 130-012, Korea 

S. F. Portegies Zwart^ 
Department of Astronomy, Boston University, 725 Commonwealth Avenue, Boston, MA 

01581, USA 

and H.A. Bethe 

Floyd R. Newman Laboratory of Nuclear Studies, Cornell University, Ithaca, New York 

14853, USA 

ABSTRACT 

At least one, but more likely two or more, eccentric neutron-star, carbon- 
oxygen white-dwarf binaries with an unrecycled pulsar have been observed. 
According to the standard scenario for evolving neutron stars which are recycled 
in common envelope evolution we expect to observe ^ 50 such circular neutron 
star-carbon oxygen white dwarf binaries, since their formation rate is roughly 
equal to that of the eccentric binaries and the time over which they can be 
observed is two orders of magnitude longer, as we shall outline. We observe 
at most one or two such circular binaries and from that we conclude that the 
standard scenario must be revised. 

Introducing hypercritical accretion into common envelope evolution removes 
the discrepancy by converting the neutron star into a black hole which does not 
emit radio waves, and therefore would not be observed. 



Subject headings: binaries: close - stars: neutron - white dwarfs - stars: 
evolution - stars: statistics 



iSPZ is Hubble Fellow 



- 2 - 



1. Introduction 

We consider the evolution of neutron-star, carbon-oxygen white-dwarf binaries using 
both the Bethe & Brown (1998) schematic analytic evolutions and the Portegies Zwart & 
Yungelson (1998) numerical population syntheses. 

The scenario in which the circular neutron-star, carbon-oxygen white-dwarf binaries 
(which we denote as {ns, co)c hereafter) have gone through common envelope evolution is 
considered. In conventional common envelope evolution for the circular binaries it is easy to 
see that the observed ratio of these to eccentric binaries (hereafter (ns, co)^:) should be ~ 50 
because: (i) The formation rate of the two types of binaries is, within a factor 2, the same, 
(ii) The magnetic fields in the circular binaries will be brought down by a factor of ~ 100 
by He accretion in the neutron-star. He-star phase following common envelope evolution 
just as the inferred pulsar magnetic field strengths in the double neutron star binaries are 
brought down (Brown 1995). In the eccentric binaries the neutron star is formed last, after 
the white dwarf, so there is nothing to circularize its orbit. More important, its magnetic 
field will behave like that of a single star and will not be brought down from the B ~ 10^^ 
G with which it is born. (At least empirically, neutron star magnetic fields are brought 
down only in binaries, by accreting matter from the companion star, Taam & Van den 
Heuvel 1986, although Wijers 1997 shows the situation to be more complex.) Neutron stars 
with higher magnetic fields can be observed only for shorter times, because of more rapid 
spin down from magnetic dipole radiation. The time of possible observation goes inversely 
with the magnetic field B. We use the observability premium 

U = 10^^G/B (1) 

(Wettig Sz Brown 1996) which gives the relative time a neutron star can be observed. Given 
our above point (ii), the circular binaries should have an observability premium H ~ 100 
as compared with H ~ 1 for the higher magnetic field neutron star in an eccentric orbit. 
Correcting for the factor 2 higher formation rate of the eccentric binaries (point (i) above) 
this predicts the factor ~ 50 ratio of circular to eccentric binaries. 

In our paper we cite one firm eccentric neutron-star, carbon-oxygen white-dwarf binary 
{ns,co)£ B2303+46 and argue for a recently observed second one, J1141— 65. Portegies 
Zwart &; Yungelson (1999) suggest PSR 1820—11 may also be in this class, but cannot 
exclude the possibility that the neutron star companion is a main sequence star (Phinney 
& Verbunt 1991). This would imply that ^ 100 such binaries with circular orbits should 
be observed. But, in fact, only one B0655+64 is observed if we accept the developing 



^ We have a special scenario for evolving it; see section 



3.2 



- 3 - 



concensus (Section p73|) that those observed (ns, co)c are evolved with avoidence of common 
envelope evolution. We are thus confronted by a big discrepancy, for which we suggest a 
solution. 

In order to understand our solution, we need to review three past works. In the earlier 
literature the observed circular {ns,co)c were evolved through common envelope, e.g., 
see Van den Heuvel (1994) and Phinney & Kulkarni (1994). Accretion from the evolving 
giant progenitor of the white dwarf was neglected, since it was thought that the accretion 
would be held to the Eddington rate of MEdd ~ 1-5 x 10~^ Mq yr~^, and in the ~ 1 year 
long common envelope evolution a negligible amount of matter would be accreted. We 
term this the standard scenario. However, Chevalier (1993) showed that once M exceeded 
~ 10'' MEdd, it was no longer held up by the radiative pressure due to the X rays from the 
neutron star, but that it swept them inwards in an adiabatic inflow. Bethe & Brown (1998) 
employed this hypercritical accretion in their evolution of double neutron star {ns, ns) and 
neutron-star, low-mass black-hole binaries {ns, Imbh) and we shall use the same techniques 
in binary evolution here. In particular, these authors found that including hypercritical 
accretion in the standard scenario for double-neutron star binary evolution, the first born 
neutron star went into a low-mass black hole. To avoid the neutron star going through 
the companion's envelope, a new scenario was introduced beginning with a double He star 
binary. It gives about the right number of double neutron star binaries. 

A new development has been that most of the circular {ns, co)c are currently evolved 
with avoidance of common envelope evolution. In Section |2]^, we shall summarize this 
work, carried out independently by King & Ritter (1999) and Tauris, Van den Heuvel & 
Savonije (2000). If we accept the new scenario, at most one or two circular {ns,co)c that 
went through common envelope evolution have been observed. Yet, in the standard scenario 
at least ~ 50 of them should be seen. 

In this paper we find that in {ns, co) which do go through common envelope evolution, 
the neutron star goes into a black hole. The {ns, co) binaries observed to date have been 
identified through radio emission from the neutron star. Thus, binaries containing a 
low-mass black hole would not have been seen. We discuss masses for which neutron stars 
go into black holes. 

Although the main point of our paper relies only on relative formation rates, we shall 
show in Appendix A that the Bethe & Brown (1998) schematic, analytic analysis agrees 
well with the detailed numerical population synthesis of Portegies Zwart, once both are 
normalized to the same supernova rate. 



-4- 



2. The Problem 
2.1. Standard Scenario vs Observations 



Portegies Zwart & Yungelson (1998), in a very careful population synthesis, have 
calculated the expected number of newly created binaries of compact stars (neutron stars or 
black holes) and white dwarfs. Among the latter, they distinguish between those consisting 
of helium and those consisting of carbon-oxygen (denoted co). To make an eccentric 
binary containing a neutron star, the supernova must occur after the carbon-oxygen star 
has formed (Portegies Zwart & Yungelson 1999). To make a circular binary containing a 
neutron star it is necessary that its companion be close so that at some earlier stage in 
evolution (but after formation of the neutron star) there was mass transfer or strong tidal 
interaction, which requires the companion to (nearly) fill its Roche Lobe. 

Since the Bethe & Brown (1998) schematic calculations did not include mass exchange, 
which is very important in evolving (ns, co)^ binaries, we need the more complete 
calculations of Portegies Zwart & Yungelson, which are listed in Table We discuss this in 
more detail later. These do not include hypercritical accretion; i.e., they follow the standard 
scenario. In this case the formation ratio of {ns,co)c to (ns, co)f is 17.7/32.1 = 0.55. 
We now make the case that if the {ns, co)c were to be formed through common envelope 
evolution (Phinney & Kulkarni 1994; Van den Heuvel 1994) in the standard scenario, their 
pulsar magnetic fields would be brought down to 5 ~ 10^*^ G because of the similarity 
to binary neutron star systems in which this occurs. In detail, this results from helium 
accretion during the neutron-star. He-star stage which precedes the final binary (Brown 
1995, Wettig & Brown 1996). 

Detailed calculation of Iben & Tutukov (1993) for original donor masses 4 — 6 Mq of the 
white dwarf progenitor show that following common envelope evolution the remnant stars 
fill their Roche lobes and continue to transfer mass to their companion neutron star. These 
remnants consist of a degenerate carbon-oxygen core and an evolving envelope undergoing 
helium shell burning. The mass transfer to the neutron star is at a rate M < lO'^MEdd? 
the lower limit for hypercritical accretion, so it is limited by Eddington. Van den Heuvel 
(1994) estimates that the neutron star accretes about 0.045 Mq and 0.024 Mq in the case 
of the ZAMS 5 Mq and 6 Mq stars, and 0.014 Mq for a 4 M© star, where these ZAMS 
masses refer to the progenitors of the white dwarfs. The accretion here is of the same order, 
roughly double,^ the wind accretion used by Wettig & Brown (1996) in the evolution of 



The He burning time to be used for the progenitor of the white dwarf is ^ 10^ years, whereas for the 



- 5 - 



the relativistic binary pulsars B1534+12 and B1913+16. There the magnetic fields were 
brought down by a factor ~ 100 from B ~ 10^^ G to ~ 10^° G, increasing the observability 
premium 11 by a factor of ~ 100. Thus, the scenario in which the {ns, co)c are produced 
through common envelope evolution without hypercritical accretion should furnish them 
with n ~ 100, by helium accretion following the common envelope. Although the detailed 
description may not be correct, the similarity of evolution of (ns, co)c with that of binary 
neutron stars in the older works (Phinney & Kulkarni 1994; Van den Heuvel 1994) should 
furnish these with about the same 11. 

There is one confirmed {ns, co)^, namely B2303+46, see Table 0, so there should be 
about 50 circular ones which went through common envelope evolution. Indeed, several 
circular ones have been observed (see Table 0), and one or two of these may have gone 
through common envelope evolution. Thus we have a big discrepancy between the standard 
scenario and the observations. In the next section, we discuss the possibility of PSR 
J1141— 6545 being {ns,co)£, which enhances the discrepancy. 

2.2. Is PSR J1141-6545 {ns,co)£ ? 

Not only is the eccentric B2303+46 quite certain, but a relativistic counterpart, PSR 
J1141— 6545 has recently been observed (Kaspi et al. 2000), in an eccentric orbit. The 
inferred magnetic dipole strength is 1.3 x 10^^ G, and the total mass is 2.300 ± 0.012 Mq. 
Kaspi et al. argue that the companion of the neutron star can only be a white dwarf, or 
neutron star. With a total mass of 2.3 Mq, if J1141— 65 were to contain two neutron stars, 
each would have to have a mass of ~ 1.15 Mq, well below the 19 accurately measured 
neutron star masses, see Fig. |I] (Thorsett & Chakrabarty 1999). 

We can understand the absence of binary neutron stars with masses below ~ 1.3 Mq, 
although neutron stars of this mass are expected to result from the relatively copious main 
sequence stars of ZAMS mass ~ 10 — 13 Mq from the argument of Brown (1997). The 
He stars in the progenitor He-star, pulsar binary of mass ^ 4 Mq (Habets 1986) expand 
substantially during He shell burning. Accretion onto the nearby pulsar sends it into a 
black hole. Indeed, with inclusion of mass loss by helium wind. He stars of masses up 
to 6 or 7 Mq expand in this stage (Woosley, Langer & Weaver 1995). Fryer & Kalogera 
(1997) find that special kick velocities need to be selected in order to avoid the evolution 
of PSR 1913+16 and PSR 1534+12 from going into a black hole by reverse Case C mass 



relativistic binary pulsars the average time of 5 x 10^ years is more appropriate, so one would expect a factor 
~ 2 greater accretion. 



-6- 



transfer (mass transfer from the evolving He star companion onto the pulsar in the He-star, 
neutron-star stage which precedes that of the binary of compact objects). 

Our above argument says that the first neutron star formed in these would be sent 
into a black hole when its companion He star evolved and poured mass on it. Therefore, we 
believe the companion in J1141— 65 must be a white dwarf. Earlier Tauris & Sennels (2000) 
developed the case that J1141— 65 was an eccentric neutron-star, white-dwarf binary. Given 
the high magnetic field of J1141— 65 (1.3 x 10^^ G) with low observability premium of 0.77, 
this would increase the predicted observed number of circular (ns, co)c which had gone 
through common envelope evolution to ~ 130 in the standard scenario. 

2.3. Evolution of Neutron-Star, Carbon-Oxygen White-Dwarf Biuciries with 

Avoidance of Common Envelope Evolution 

Our discussion of the common envelope evolution in the last section applied to 
convective donors. In case the donor is radiative or semiconvcctive, common envelope 
evolution can be avoided. Starting from the work of Savonijc (1983), Van den Heuvel 
(1995) proposed that most low mass X-ray binaries would evolve through a Her X-1 type 
scenario, where the radiative donor, more massive than the neutron star, poured matter 
onto its accretion disk at a super Eddington rate, during which time almost all of the 
matter was flung off. This involved Roche Lobe overflow. Although Van den Heuvel hmited 
the ZAMS mass of the radiative donor to 2.25 Mq in order to evolve helium white-dwarf, 
neutron star binaries, his scenario has been extended to higher ZAMS mass donors in order 
to evolve the carbon-oxygen white-dwarf, neutron star binaries. The advection dominated 
inflow-outflow solutions (ADIOS) of Blandford & Bcgclman (1999) suggest that the binding 
energy released at the neutron star can carry away mass, angular momentum and energy 
from the gas accreting onto the accretion disk provided the latter does not cool too much. 
In this way the binding energy of gas at the neutron star can carry off ~ 10^ grams of gas at 
the accretion disk for each gram accreting onto the neutron star. King & Begelman (1999) 
suggest that such radiatively-driven outflows allow the binary to avoid common envelope 
evolution. 

As noted above, for helium white dwarf companions. Van den Heuvel (1995) had 
suggested Cyg X-2 as an example following the Her X-1 scenario. King & Ritter (1999) 
calculated the evolution of Cyg X-2 in the ADIOS scenario in detail. These authors also 

evolved the {ns,co)c binaries in this way, using donor stars of ZAMS masses 4 — 7 Mq. 
Tauris, Van den Heuvel, & Savonije (1999) have carried out similar calculations, with stable 
mass transfer. These authors flnd that even for extremely high mass-transfer rates, up to 



- 7- 



M ~ lO^MEdd, the system will be able to avoid a common envelope and spiral-in evolution. 

Tauris, van den Heuvel & Savonije 2000 evolve J1453-58, J1435-60 and J1756-5322, 
the three lowest entries in our Table H, through common envelope. We obtained the 
eccentricities and P's for the first two of these (Fernando Camilo, private communication). 
The binary J1453— 58, quite similar to J062H-1002 has a substantial eccentricity and clearly 
should be evolved with a convective donor as Tauris et al did for J0621+1002. The spin 
periods of J1435— 60 and J1756— 5322 are short, indicating greater recycling than the other 
listed pulsars. It would seem difficult to get the inferred magnetic field down to the 4.7 x 10^ 
G of J1435— 60 by the Iben & Tutukov or Wettig & Brown accretion scenarios following 
common envelope evolution as discussed in Section |2.1| . If, however, one does believe that 
J1435— 60 and J1756— 5322 have gone through common envelope, the discrepancy between 
predicted and observed circular binaries in the standard scenario is only slightly relieved. 



2.4. Are There Observational Selection Effects ? 

In Table ^ we have tabulated 5*400 ^ in order to see whether the normalized intensity 
gives strong selection effects. Note that the 35.95 for B2303-I-46 is not so different from 
the 43.56 and 203.35 for B1534+12 and B1913+16, respectively. For the circular binaries 
(ns, co)c the intensities are less, but their empirical Observability Premium 11 is much 
larger. There may be other observational selection effects, but, we believe that there are no 
observational selection effects strong enough to compensate for the factor 100 discrepancy 
between the observed population and the one expected from the standard model. So the 
problem remains the same. 



3. The Answer 

3.1. Black Hole Formation in Common Envelope Evolution 

We believe the answer to the missing binaries is that the neutron star goes into a black 
hole in common envelope evolution, as we now describe. We label the mass of the neutron 
star as Ma and that of the giant progenitor of the white dwarf as Mb- Following Bethe & 
Brown (1998) we choose as variables the neutron star mass Ma and Y = Mb /a, where a is 
the orbital radius. From their eq. (5.12) we find 



Maj ^ /Yf_- 
Ma, Vr. 



(2) 



- 8 - 



where q is the drag coefficient. From Shima et al. (1985) we take 

Q = 6. (3) 

We furnish the energy to remove the hydrogen envelope of the giant B (multiphed by ot^}, 
where ace is the efficiency of couphng of the orbital motion of the neutron star to the 
envelope of B) by the drop in orbital energy of the neutron star; i.e., 



Here the CGGMB^jFj is just the binding energy of the initial giant envelope, found by 
Applegate (1997) to be CGGM^^^a,"^, and the right hand side of the equation is the final 
gravitational binding energy ^GM^jMBjaJ^ in our variables. Using eqs. (|^) and (^ in 
eq. (^ one finds 

l/Cd 

(5) 



Maj _ (l.2MB,i 



MA,i \aceMA,i) 

For the sake of argument, we take the possible range of initial neutron star mass to be 
1.2 — 1.5 Mq (the upper bound is the Brown & Bethe (1994) mass at which a neutron 
star goes into a low-mass black hole), and the main sequence progenitor masses of the 
carbon-oxygen white dwarf to be MB,i = 2.25 — 10 Mq. As we show in Appendix C, in the 
Bethe & Brown (1998) schematic model, mass transfer was assumed to take place when the 
evolving giant reached the neutron star, whereas more correctly it begins when the envelope 
of the giant comes to its Roche Lobe. For the masses we employ, main sequence progenitors 
of the carbon-oxygen white dwarf of 2.25 — 10 Mq, the fractional Roche Lobe radius is 

tl ~ 0.5. (6) 

The binding energy of the progenitor giant at its Roche Lobe is, thus, double what it would 
be at Oi, the separation of giant and neutron star. Therefore, a Bethe & Brown ace = 0.5 
corresponds to a true efficiency dee ~ 1, if the latter is defined as the value for which the 
envelope removal energy, at its Roche Lobe, is equal to the drop in neutron star orbital 
energy as it moves from ai to a/. If we take ace = 0.5 in eq. (^ we find, given our assumed 
possible intervals 

1.54 Mq ^ Maj ^ 2.38 Mq. (7) 

These are above the neutron star mass limit 1.5 Mq (Brown & Bethe 1994) beyond which a 
neutron star goes into a low-mass black hole. Thus, all neutron stars with common envelope 
evolution in our scenario evolve into black holes. This solves the big discrepancy between 
the standard scenario and observation. The only remaining problem is the evolution 
of B0655+64, which survived the common envelope evolution, and we suggest a special 
scenario for it in the next section. 



-9- 



3.2. Is B0655+64 a problem? 

Van den Heuvel & Taam (1984) were the first to notice that the (ns,co)c system 
B0655+64 might have been formed in a similar way as the double neutron stars. The short 
period of 1.03 days, magnetic field ~ 10^*^ G, and the high companion mass of ~ 1 Mq make 
this binary most similar to a binary neutron star, but with a carbon-oxygen white-dwarf 
companion, resulting from probable ZAMS masses ~ 5 — 8 Mq. For a 1.4 Mq neutron star 
with 1 Mq white-dwarf companion af — 5.7 Rq. 

The similarity of B0655-I-64 to the close neutron star binaries suggests the double 
hehum star scenario (Brown 1995) to calculate the evolution. The ZAMS mass of the 
primary is chosen to be just above the limit for going into a neutron star, that of the 
secondary just below. For the double He star scenario the ZAMS masses of primary and 
secondary cannot be more than ~ 5% different. However, in this case the ratio q of masses 
is so close to unity that the secondary will not be rejuvenated (Braun & Langer 1995: If the 
core burning of hydrogen to helium in the companion star is nearly complete, the accreted 
matter would have to cross a molecular weight barrier in order to burn and if q is near 
unity there is not time enough to do so. Thus He cores of both stars will evolve as if the 
progenitors never had more than their initial ZAMS mass.) 

What we have learned recently about effects of mass loss (Wellstein & Langer 1999) will 
change the Brown (1995) scenario in detail, but not in general concept. An ~ 10 Mq ZAMS 
star which loses mass in RLOF to a lower mass companion will burn helium as a lower-mass 
star due to subsequent mass loss by helium winds, roughly as an 8 Mq star (Wellstein & 
Langer, in preparation). Thus, the primary must have ZAMS mass ^ 10 Mq in this case 
in order to evolve into a neutron star following mass loss. Although the secondary will not 
be rejuvenated as mass is transferred to it, it will burn helium without helium wind loss 
because it is clothed with a hydrogen envelope. Thus, a secondary of ZAMS 8 Mq will burn 
He roughly as the primary of 10 Mq in the situation considered. Given these estimates, a 
primary of ZAMS mass M ^ 10 Mq will evolve into a white dwarf, whereas a secondary of 
mass ^ 8 Mq will end up as a neutron star. Of course, the former must be more massive 
than the latter, but stars in this mass range are copious because this is the lowest mass 
range from which neutron stars can be evolved, so there will be many such cases. 

This scenario might not be as special as outlined because the fate of stars of ZAMS 
mass 8 — 10 Mq, which do not form iron cores but do burn in quite different ways from 
more massive stars, is somewhat uncertain in the hterature. Whereas it is generally thought 
that single stars in this range end up as neutron stars, it has also been suggested that some 
of them evolve as AGB stars ending in white dwarfs. In terms of these discussions it does 
not seem unlikely that with two stars in the binary of roughly the same mass, the first 



- 10 - 



to evolve will end up as a neutron star and the second as a white dwarf, especially if the 
matter transferred in RLOF cannot rejuvenate the companion. 

Van den Heuvel & Taam (1984) evolved B0655+64 by common envelope evolution. 
In taking up the problem again, Tauris, Van den Heuvel & Savonije (2000) in agreement 
with King & Ritter (1999) find that B0655+64 cannot be satisfactorily evolved with their 
convective donor scenario. Tauris et al. suggest a spiral-in phase is the most plausible 
scenario for the formation of this system, but we find that the neutron star would go into a 
black hole in this scenario, unless the two progenitors burn He at the same time. 



3.3. Neutron Star Masses 

There is by no means agreement about maximum and minimum neutron star masses 
in the literature. The mass determination of Vela X-1 have been consistently higher than 
the Brown & Bethe 1.5 Mq which is consistent with well measured neutron star masses in 
Fig. |I|. In a recent careful study at ESO Barziv et al. (2000), as reported by Van Kerkwijk 
(2000), obtain 

M^s = 1.87^:°:?? Mq. (8) 

Even at 99% confidence level, M7V5 > 1.6 Mq. Taking the maximum mass to be 1.87 Mq 
and ace = 0.5, {Miqs)min = 1-2 Mq one finds from eq. (^ that the maximum carbon-oxygen 
white dwarf progenitor mass of (ns, co)c is 



Although there is some uncertainty in the efficiency a^ei the ratio Ms^i/ Mq is much more 
sensitive to Maj because of the 6th power of the ratio in eq. (^).0 But then one cannot 
explain why no (ns, co)c (except B0655+64) which survived the common envelope evolution 
are seen, since this mass is high enough to give those of the white dwarf companions. 

Distortion of the ~ 20 Mq B-star companion by the neutron star in Vela X-1 brings 
in large corrections (Zuiderwijk et al. 1977, van Paradijs et al. 1977a, van Paradijs et al. 
1997b) making measurement of neutron star masses in high-mass X-ray binaries much more 
difficult than those with degenerate companions. 



With Mns = 1-5 Mg we get [MB,i)max = 1-9 Mq which is below the minimum Mb 2.25 Mq) for 
forming a carbon-oxygen white dwarf, so no [ns, co)c survive the common envelope evolution. 



- 11 - 



Given the ~ 0.43 at the collapse of the core of a large star (Aufderheide et al. 1990) 
one finds the cold Chandrasekhar mass to be 

Mcs = 5.76 Mq ^ 1.06 M© (10) 

where Yg is the ratio of the number of electrons to the number of nucleons. Thermal 
corrections increase this a bit, whereas lattice corrections on the electrons decrease it, so 
that when all is said and done, Mqs ^1-1 Mq (Shapiro & Teukolsky 1983). The major 
dynamical correction to this is from fallback following the supernova explosion. We believe 
that fallback in supernova explosions will add at least ~ 0.1 Mq to the neutron star, since 
bifurcation of the matter going out and in happens at about 4000 km (Bethe & Brown 
1995). Thus our lower limit of ~ 1.2 Mq is reasonable. 



4. Discussion and Conclusions 

At least one, but more likely two or more, {ns, co)^ binaries with an unrecycled pulsar 
have been observed. According to the standard scenario for evolving neutron stars which 
are recycled in a common envelope evolution we then expect to observe ^ 50 (ns, co)c- We 
only observe B0655+64 (which we evolve in our double He-star way) and possibly one or 
two binaries that went through common envelope evolution and from that we conclude that 
the standard scenario must be revised. Introducing hypercritical accretion into common 
envelope evolution (Brown 1995; Bethe & Brown 1998) removes the discrepancy. 

We believe that the evolution of the other (ns, co)c binaries may originate from systems 
with a neutron star with a radiative or semi-convective companion. The accretion rate in 
these systems can be as high as lO^'MEdd but common envelope evolution is avoided. This 
possibility, however, does not affect our conclusion concerning hypercritical accretion. 

It is difficult to see "fresh" (unrecycled) neutron stars in binaries because they don't 
shine for long. B2303+46 (Table ^ is the most firm example of a {ns,co)£ binary with 
a fresh neutron star. Although binaries where a "fresh" neutron star is accompanied by 
a black hole have similar birthrates (~ 10~^ yr~^ for both types; Bethe & Brown, 1999, 
and Portegies Zwart & Yungelson 1999) and lifetime, none are observed. In Appendix 
A we quote results of Ramachandran & Portegies Zwart (1998) that show there is an 
observational penalty which disfavors the observation of neutron stars with black holes as 
companions, because of the difficulty in identifying the pulsar due to the Doppler shift 
which smears out the signal in these short-period objects. Because of the longer orbital 
period and lower companion mass of (ns, co)£, such binaries are less severely plagued by 
this effect, although the recently discovered J1141— 6545 is a relativistic binary with 5 hr 



- 12 - 



period. We therefore argue that it is not unreasonable that no {Imbh, ns) binaries have yet 
been observed, but that they should be actively searched for since the probability of seeing 
them is not far down from that of seeing the (ns, co)£^s. 

We are grateful to Fernando Camilo for information in Table ^ We would like to thank 
Brad Hansen, Marten van Kerkwijk, Ralph Wijers, Thomas Tauris and Lev Yungelson for 
useful discussions and advice, and thank Justin Holmer for providing us with the convective 
envelopes for stars in the giant phase. GEB would like to thank Brad Hansen also for 
correspondence which started him on this problem. GEB & CHL were supported by the 
U.S. Department of Energy under grant DE-FG02-88ER40388. 

SPZ was supported by NASA through Hubble Fellowship grant HF-01112.01-98A by 
the Space Telescope Science Institute, which is operated by the Association of Universities 
for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. SPZ is grateful 
for the hospitality of the State University of New York at Stony Brook. 

A. Comparison of Population Syntheses 

In this Appendix we first compare results that the Bethe & Brown (1998) schematic 
analytic evolution would have given without hypercritical accretion with the Portegies 
Zwart & Yungelson (1998, 1999) results of Table |1], which do not include hypercritical 
accretion. We can then illustrate how hypercritical accretion changes the results. 

Without hypercritical accretion the {Imbh, ns) binaries of Bethe & Brown would end 
up rather as neutron star binaries {ns,ns), giving a summed formation rate of 1.1 x 10~^ 
yr~^, to compare with 1.1 x 10~^ yr~^ from the Portegies Zwart numerical driven population 
sythesis results presented in Table |l|. This good agreement indicates that kicks that the 
neutron star receives in formation were implemented in the same way in the two syntheses. 
Introduction of hypercritical accretion leaves only those neutron stars which do not go 
through a common envelope; i.e., those in the double He star scenario of Brown (1995), with 
formation rate 10~^ yr~^. This is much closer to the estimated empirical rate of 8 x 10"^ 
yr~^ of Van den Heuvel & Lorimer (1996) which equals the rate derived independently by 
Narayan et al. (1991) and Phinney (1991). Large poorly known factors are introduced in 
arriving at these "empirical" figures, so it is useful that our theoretical estimates end up 
close to them. In our theoretical estimates the possibility described earlier that the pulsar 
in the lower mass binary pulsars goes into a black hole in the He shell burning stage of 
the progenitor He-star, neutron-star binary (Brown 1997) was not taken into account and 
this process may change ~ half of the remaining neutron star binaries in our evolution into 



- 13 - 



(Imhh^ns) binaries. 

Bethe & Brown (1998) had a numerical symmetry between high-mass binaries in which 
both massive stars go supernova and those in which the more massive one goes supernova 
and the other, below the assumed dividing mass of 10 M©, did not; i.e., the number of 
binaries was equal in the two cases. Taking ZAMS mass progenitors of 2.3 — 10 Mq for 
carbon-oxygen white dwarfs, we then find a rate of 

10 — 2 3 

R = 2x — X 1.1 X 10"^ yr"^ = 16.9 x 10"^ yr"^ (Al) 

for the formation rate of {ns,co)c binaries. The 1.1 x 10^"^ yr~^ is taken from the last 
paragraph and applies here because of the numerical symmetry mentioned above. The 
factor 2 results because there is no final explosion of the white dwarf to disrupt the binary, 
as there was above in the formation of the neutron star. This rate R is to be compared with 
(ns, co)c = 17.7 X 10~^ yr~^ in Table |I[ These (ns, co)c binaries are just the ones in which 
the neutron star goes into a black hole in common envelope evolution, unless the masses 
of the two initial progenitors are so close that they burn He at the same time. Then a 
binary such as B0655+64 can result, since the two helium stars then go through a common 
envelope, rather than the neutron star and main sequence star. 

It is of interest to compare the populations of {ns, ns) binaries with the {ns, co)^ 
binaries. We must rely on the Portegies Zwart result for the latter, which cannot be evolved 
without mass transfer, which is not included in the Bethe & Brown evolution. The {ns, ns) 
binaries involve common envelope evolution, whereas the {ns, co)£ do not. Thus, results 
for the rates should differ substantially in the standard scenario, which does not include 
hypercritical accretion, and our scenario which does. The ratio for {ns,ns) and {ns,co)£ 
from Table [l| are 10.6 x 10~^ yr~^ and 32.1 x 10~^ yr~^- The {ns,ns) are recycled in the 
He-star, pulsar stage by the He wind, giving an observability premium of H ~ 100 (Brown 
1995). The pulsar in the {ns,co)s is not recycled. Thus, the expected observational ratio is 

100X10.6^33 1^2^ 



{ns,co)£ 32.1 

Now B2303+46, and possibly B1820— 11 and J1141— 6545 lie in the {ns,co)£ class, whereas 
B1534-M2 and B1913+16 are relativistic binary neutron stars with recycled pulsars. We do 
not include the neutron star binary 2127+llC, although it has the same B as the other two. 
It is naturally explained as resulting from an exchange reaction between a neutron star and 
a binary which took place < 10^ years ago in the cluster core of M15 (Phinney & Sigurdsson 
1991). Thus, the empirical ratio eq. (|A2|) is not much different from unity. In Bethe & 
Brown (1998) common envelope evolution cuts the {ns,ns) rate down by a factor of 11, 
only the 1/11 of the binaries which burn He at the same time surviving. The remaining 



-14- 



factor 3 is much closer to observation. Furthermore, Ramachandran & Portegies Zwart 
(1998) point out that there is an observational penalty of a factor of several disfavoring 
the relativistic binary neutron stars because of the difficulty in identifying them due to the 
Doppler shift which smears out the signal in these short-period objects. Some observational 
penalty should, however, also be applied to J1141— 6545, which is a relativistic binary. We 
estimate that the combination of neutron stars going into black holes in common envelope 
evolution and the greater difficulty in seeing them will bring the ratio of 33 in eq. ( [^) 
down to ~ 1 or 2, close to observation. 

We have shown that there is remarkable agreement between the Bethe & Brown (1998) 
schematic analytic population synthesis and the computer driven numerical synthesis of 
Portegies Zwart & Yungelson (1998). This agreement can be understood by the scale 
invariance in the assumed logarithmic distribution of binary separations. In general we are 
interested in the fraction of binaries which end up in a given interval of a. E.g., in Bethe & 
Brown (1998) that fraction was 

# = ^ (A3) 

where d{\\ia) was the logarithmic interval between the Oj below which the star in the binary 
would merge in common envelope evolution and a/, the largest radius for which they would 
merge in a Hubble time. Here 7 was the assumed initial logarithmic interval over which the 
binaries were distributed. Thus, the desired fraction 

rf(lna) = Aa/a (A4) 

is scale invariant. Mass exchange in the evolution of the binary will change the values of 
Oj and Qf delineating the favorable logarithmic interval, but will not change the favorable 
c?(lna). Of course, when He stars go into neutron stars, the probability of the binary 
surviving the kick velocity does depend on the actual value of a, violating the scale 
invariance. But this does not seem to be a large effect in the calculations. In the case of 
the formation of (ns, co)s binaries, the neutron star is formed last, out of the more massive 
progenitor. Mass transfer is required for this, because otherwise the more massive progenitor 
would explode first. The mass must not only be transferred, but must be accepted, so 
that the companion star is rejuvenated (unless g ~ 1 as discussed). We need the Portegies 
Zwart & Yungelson detailed numerical program for this. In fact, in calculations with this 
program (See Table ^ the formation of (ns, co)s binaries is nearly double that of (ns, co)c 
ones. However, for q ^ 0.75, where q is the mass ratio of original progenitors, of the ZAMS 
progenitors Braun & Langer (1995) showed that the transferred hydrogen has trouble 
passing the molecular weight barrier in the companion, so that the latter would not be 
rejuvenated. We have not included this effect here, but roughly estimate that it will lower 



- 15 - 



the predicted numbers of {ns, co)^ by a factor > 2, bringing it down below the number of 
{ns, co)c, exacerbating the problems of the standard model of binary evolution. 

In the literature one sees statements such as "Population syntheses are plagued by 
uncertainties". It is, therefore, important to show that when the same assumptions about 
binary evolution are made and when the syntheses are normalized to the same supernova 
rates, similar results are obtained. The evolution in the Bethe & Brown (1998) schematic 
way is simple, so that effects in changes in assumptions are easily followed. 



B. Hypercritical Accretion 

We develop here a simple criterion for the presence of hypercritical accretion. We 
further show that if it holds in common envelope evolution for one separation a of the 
compact object met by the expanding red giant or supergiant, it will also hold for other 
separations and for other times during the spiral in. We assume the envelope of the giant 
to be convective. 

In the rest frame of the compact object, Bondi-Hoyle-Lyttleton accretion of the 
envelope matter (hydrogen) of density poo and velocity V is (for F = 5/3 matter) 

M = 2.23 X W^\MJ Mq)X~'Poo g s"^ (Bl) 

where M^o is the mass of the compact object, and Vg is the velocity in units of 1000 km s~^. 
Poo is given in g cm~^. From Brown (1995) the minimum rate for hypercritical accretion is 

=1.09x10^ (B2) 



Ml 



Edd 



For hydrogen 

Mcr = 0.99 X lO^^g s-\ (B3) 
Using eqs.([BTD and ( [B2D we obtain 

(Poo)cr = 0.44 X 10-^( MQ/M,,fVi g cm-\ (B4) 
Using Kepler for circular orbits 

V' = (B5) 



16 



where M^ot is the mass of the compact object plus the mass of the hehum core of the 
companion plus the envelope mass interior to the orbit of the compact object. One finds 



The a-dependence of (poo)cr is the same as the asymptotic density for the n = 3/2 polytrope 
which describes the convective envelope. Thus, if the criterion for hypercritical accretion is 
satisfied at one time and at one radius it will tend to be satisfied for other times and for 
other radii. The change of Mfot with a is unimportant because from Table |^ it can be seen 
that p > (poo)cr already near the surface of the star. 

In order to check the applicability of hypercritical accretion to the compact object in 
the relatively low-mass stars we consider in this paper, we make application to a 4 Mq red 
giant of radius R = 100 Rq, evolved as pure hydrogen but with inclusion of dissociation 
by Justin Holmer (1998). In the Table ^ we compare the densities in the outer part of the 
hydrogen envelope with those needed for hypercritical accretion. From the table it can 
be seen that hypercritical accretion sets in quickly, once the compact object enters the 
envelope of the evolving giant. 

Note that the accretion through most of the envelope will be > 1 Mq yr~^. Since the 
total mass accreted by the neutron star is ~ 1 Mq this gives a dynamical time of ^ 1 year, 
although the major part of the accretion takes place in less time. This is in agreement with 
the dynamical time found, without inclusion of accretion, by Terman, Taam & Hernquist 



We discuss the definition of the efficiency of the hydro dynamical coupling of the orbital 
motion of the neutron star to the envelope of the main sequence star. 

Van den Heuvel (1994) starts from the Webbink (1984) energetics in which the 
gravitational binding energy of the hydrogen envelope of the giant is taken to be 




(B6) 



(1995). 



C. EfRciency 



E, 



'env 



G(Mco,e + Menv)M, 

R 



env 



(CI) 



which results in the envelope gravitational energy 



E, 



'env 



0.7GM^ 
R 



(C2) 



-17- 



where M = Mcore + M^nv is the total stellar mass and the Bethe & Brown (1998) 
approximation Mcorc — 0.3M has been used. 

Applegate (1997) has calculated the binding energy of a convective giant envelope, 
obtaining 

Eb = -O.QGMyR = ^E,nv (C3) 

Note that M is the total stellar mass, also that Eb is just 1/2 of the gravitational potential 
energy, the kinetic energy being included in Eb- Eq. ( |C3|) was checked independently by 
Holmer (1998). Unfortunately, this work was never published. Van den Heuvel and others 
have introduced an additional parameter A that both takes into account the kinetic energy 
and the density distribution of the star, R in eqs. ( |CT| ) & ( |C^ ) being replaced by RX. They 
use 



With A = 7/6 this is the same as Eb in eq. ( |C3|) . Van den Heuvel (1994) chooses A = 1/2. 
The result is that his efficiency t] is a factor of 7/3 too high. Thus, his suggested efficiency 
?7 = 4 is more like 77 ~ 12/7 = dee- 

Bethe & Brown (1998) used the Applegate result but incorrectly took the necessary 
energy to expel the giant envelope as 

Eg = -0.6 GM^ai (C5) 

rather than 

Eg = -0.6 GM^aiTL, (C6) 

the latter being the correct energy needed to remove the giant envelope at its Roche Lobe. 
The correct efficiency is 

ace = {ace)BB/rL- (C7) 

and since for the binaries considered here with g ~ 4 the fractional Roche Lobe is tl ~ 0.5, 

dee ^ 2(ace)BB = 1, (C8) 
with the Bethe & Brown (1998) a^e = 0.5. 

Of course, dee should not vary with Roche Lobe, the Bethe & Brown (1998) usage of 
Oce being in error. 



-18- 



For OLce — 1, the envelope would be removed from the giant, but would end up with 
zero kinetic energy, which is unreasonable. Thus, without additional energy soTirces, as 
discussed earlier in our note, one would expect dice ~ 0.5, in which case the kinetic energy of 
the envelope would remain unchanged in its expulsion. The Bethe & Brown (1998) results 
were insensitive to changes in dee, which changed the location but not the magnitude of the 
favored logarithmic intervals, as noted by those authors. 

In the present case Van den Hcuvcl's Ace = 1.2 definitely indicated the presence of 
energy sources additional to the drop in orbital energy, although they arc not as large as he 
indicated. We have checked that with a^e = 1.2 and q ^ 1 in the Bethe & Brown (1998) 
formation we obtain the numerical results of Table 1 of Van den Heuvel (1994). 

REFERENCES 
Applegate, J.H. 1997, Columbia Univ. Preprint 

Aufderheide, M.B., Brown, G.E., Kuo, T.T.S., Stout, D.B., & Vogel, P. 1990, ApJ, 362, 241 

Bailes, M., Harrison, P.A., Lorimer, D.R., Johnson, S., Lyne, A.G., Manchester, R.N., 
D'amico, N., Nicastro, L., Tauris, T.M., & Robinson, C. 1994, ApJ, 425, L41 

Barziv, O., Kaper, L., Van Kerkwijk, M.H., Telting, J., & Van Paradijs, J., 2000, in 
preparation 

Bethe, H. A., & Brown, G. E. 1995, ApJ, 445, L129 

Bethe, H. A., & Brown, G. E. 1998, ApJ, 506, 780 

Bhattacharya, D., Van den Heuvel, E.P.J. 1991, Phys. Rep., 203, 1 

Blandford, R.D., & Begelman, M.C. 1999, MNRAS 303, LI 

Braun, H., & Langer, N. 1995, A&A, 297, 483 

Brown, G. E. 1995, ApJ, 440, 270 

Brown, G. E. 1997, Physikahsche Blaetter, 53, 671 

Brown, G. E., & Bethe, H.A. 1994, ApJ, 423, 659 

Camilo, F. 1995, Princeton University Thesis: ''A Search for Millisecond Pulsars" 
Camilo, F., Nice, D.J., Shrauner, J.A., & Taylor, J.H. 1996, ApJ, 469, 819 



- 19 - 



Chevalier, R.A. 1993, ApJ, 411, L33 

Fryer, C.L., Benz, W., Herant, M. 1996, ApJ, 460, 801 

Fryer, C, & Kalogera, V. 1997, ApJ, 489, 244 

Habets, G.M.B.H. 1986a, A&A, 165, 95; 1986b, A&A, 167, 61 

Hansen, B. 1999, private communication 

Holmer, J. 1998, Stony Brook Senior Thesis (unpubhshed) 

Iben, I., & Tutukov, A.V. 1993, ApJ, 418, 343 

Kaspi, V.M., Lyne, A.G., Manchester, R.N., Crawford, F., Camilo, F., Bell, J.F., D'Amico, 
N., Stairs, I.H., McKay, N.P.F., Morris, D.J., & Possenti, A. 2000, ApJ, accepted; 
astro-ph/0005214l 

King, A.R., Begelman, M.C. 1999, ApJ, 519, L169 

King, A.R., Ritter, H. 1999, MNRAS, 309, 253 

Lorimer, D.R., Lyne, A.G., Bailes, M., Manchester, R.N., D'amico, N., Stappers, B.W., 
Johnston, S., & Camilo, F. 1996, MNRAS, 283, 1383 

Narayan, R., Piran, T., Shemi, A., 1991, ApJ, 379, L17 

Phinney, E.S. 1991, ApJ, 380, L17 

Phinney, E.S., Kulkarni, S.R. 1994, ARA&A, 32, 591 

Phinney, E.S., Sigurdsson, S. 1991, Nature 349, 220 

Phinney, E.S., & Verbunt, F. 1991, MNRAS, 248, 21 

Portegies Zwart, S.F., & Yungelson, L.R. 1998, A&A, 332, 173 

Portegies Zwart, S.F., & Yungelson, L.R. 1999, MNRAS 309, 26 

Ramachandran, R., & Portegies Zwart, S.F. 1998, Abstracts of the 19th Texas Symposium 
on Relativistic Astrophysics and Cosmology, Paris, France, Dec. 14-18, 1998, eds. J. 
Paul, T. Montmerle, and E. Aubourg 

Savonije, G.J. 1983, Nature, 304, 422 

Schaller, G., Schaerer, D., Meynet, G., & Maeder, A. 1992, A&AS, 96, 269 



- 20 - 



Shapiro, S.L. and Teukolsky 1983, "Black Holes, White Dwarfs, and Neutron Stars : The 
Physics of Compacy Objects" , John Wiley & Sons, Inc. 

Shima, E., Matsuda, T., Takeda, H., & Sawada, K. 1985, MNRAS, 217, 367 

Taam, R.E. and Van den Heuvel, E.P.J. 1986, ApJ, 305, 235 

Tauris, T.M., Sennels, T. 2000, A&A, 355, 236 

Tauris, T.M., Van den Heuvel, E.P.J., Savonije, G.J. 2000, ApJ, 530, L93 

Terman, J.L., Taam, R.E., & Hernquist, L. 1995, ApJ, 445, 376 

Thorsett, S.E., & Chakrabarty D. 1999, ApJ, 512, 288 

Van den Heuvel, E.P.J. 1994, A&A, 291, L39 

Van den Heuvel, E.P.J. 1995, J. Astrophys., 16, 255 

Van den Heuvel, E.P.J. , & Lorimer, D.R. 1996, MNRAS, 283, L37 

Van den Heuvel, E.P.J. , & Taam, R.E. 1984, Nature, 309, 235 

Van Kerkwijk 2000, Proceedings of ESO Workshop on "Black Holes in Binaries and Galactic 
Nuclei", Garching, Sept. 1999, eds. L. Kaper, E.P.J, van den Heuvel, and P.A. 
Woudt, Springer- Verlag; |astro-ph/0001077| 

Van Kerkwijk, M.H., & Kulkarni, S.R. 1995, ApJ, 454, L141 

Van Kerkwijk, M.H., & Kulkarni, S.R. 1999, ApJ, 516, L25 

Van Paradijs, J., Takens, R., & Zuiderwijk, E. 1977a, A&A 57, 221 

Van Paradijs, J., Zuiderwijk, E., Takens, R.J., Hammerschlag-Hensberge, G., 1977b, A&AS 
30, 195 

Webbink, R.F. 1984, ApJ, 277, 355 

Wellstein, S. & Langer, N. 1999, A&A, 350, 148 

Wettig, T., Brown, G.E. 1996, New Astronomy, 1, 17 

Wijers, R.A.M.J. 1997, MNRAS, 287, 607 

Woosley, S.E., Langer, N., & Weaver, T.A. 1995, ApJ, 448, 315 



- 21 - 

Zuiderwijk, E.J., Hammerschlag-Hensberge, G., van Paradijs, J., Sterken, C, & Hensberge, 
H. 1977, A&A 54, 167 



This preprint was prepared with the AAS IM^jX macros v4.0. 



-22- 



Table 1: Simulations, normalized to supernova rate of 0.025 yr~^, assuming 100% binarity 
following Case B of Portegies Zwart & Yungelson (1998). These simulations do not include 
hypercritical accretion. 



binary 


model B 




[10-5 yr-i] 


{ns, ns) 


10.6 


{bh, ns) 


1.9 


{ns, co)c 


17.7 


{ns, co)g 


32.1 



-23- 



Table 2: Binary Radio Pulsar Systems : {ns, ns) and {ns, co) binaries. The Observability 
Premium 11 = [lO^"^ G]/ B. Mp (Mc) means the pulsar (companion) mass, and / the mass 
function. 



IT Uiodl 


P 1, 

orb 


P . 

^ spin 


f 

/ 


M 


M 


£3 


A 

ti 


R 


n 




[uaybj 


[msj 


[ -'"0J 


L -'"0J 


L -'"0J 




[knrl 
[Kpcj 






I /to , /to I 




















Tl ^1 R -u zLQnzL 




An Q 


U. 1 ID 


\ 1.(0 


^ n Q'? 

^ u. yo 


n 9zLQ 


n 7n 








n 491 










n 97zL 

U.Z ( ^ 


1 1 

1 . 1 


lU 


1 nn 


Rl Q1 Q _U 1 f^ 


u.ozo 


oy.u 






1.00 / 


n f^i 7 

u.oi / 


/ .10 


9 V 1 n^^ 


Al 


R9197_u llPt 


u.ooo 


OyJ.o 






l.OUO 


U.Uol 


1 n 


1 9 V inio 

±.Zi A XU 


00 






















B2303 + 46 


12.34 


1066. 


0.246 


< 1.44 


> 1.20 


0.658 


4.35 


7.9 x 10^^ 


1.26 


J1141 -6545tt 


0.198 


394. 


0.177 


< 1.348 


> 0.97 


0.172 


3.2 


1.3 X 10^2 


0.77 


(ns, 00)^"^ 




















J2145 - 0750 


6.839 


16.1 


0.024 




0.515 


2.1x10-^ 


0.5 


6 X 10^ 


1667 


J1022 + 1001 


7.805 


16.5 


0.083 




0.872 


9.8x10-5 


0.6 


8.4 X 10^ 


1190 


J1603 - 7202 


6.309 


14.8 


0.009 




0.346 


< 2x10-5 


1.6 


4.6 X 10^ 


2173 


J0621 + 1002 


8.319 


28.9 


0.027 




0.540 


0.00245 


1.9 


1.6 X 10*^ 


625 


B0655 + 64 


1.029 


195.7 


0.071 




0.814 


0.75x10-^ 


0.48* 


1.26 X 10^° 


79 


J1810 - 2005 


15.01 


32.8 


0.0085 




0.34 






2.1 X 10*^ 


476 


J1157- 5112 


3.507 


43.6 


0.2546 




> 1.20 






< 6.3 X 10^ 


159 


J1232 -6501 


1.863 


88.3 


0.0014 




0.175 






9.5 X 10^ 


105 


J1453 - 58 


12.42 


45.3 


0.13 




1.07 


0.0019 




6.1 X 10^ 


164 


J1435 -60 


1.355 


9.35 


0.14 




1.10 


1x10-5 




4.7 X 10^ 


2127 


J1756 - 5322 


0.453 


8.87 


0.0475 




0.683 











f: binary in globular cluster M15. fj: not confirmed yet. assumed distance. 

irk The white dwarf mass Mc is calculated assuming Mp = 1.4 Mq and i — 60°. 

Refs; Thorsett et al. (1999); B2303: Kcrkwijk et al. (1999); J2145: Bailes et al. (1994); 

J1022: Camilo (1995); J1603: Lorimer et al. (1996); J0621: Camilo et al. (1996); B0655: 

Kerkwijk et al. (1995); J1141-65: Kaspi et al. (2000). 



-24- 



Table 3: Flux densities at 400 MHz [Ref: The Pulsar Catalog, Princeton Pulsar Group, 
http: / / pulsar .princeton.eHu 



Pulsar 


'S'400 


S'400 X d"^ 




[mJy] 


[mJy 


• kpc2] 


(ns, ns) 








51534 + 12 


36.00 




43.56 


51913 + 16 


4.00 




203.35 


52127 + lie 


0.60 




56.45 


(ns, co)s 








52303 + 46 


1.90 




35.95 


(ns, co)c 








J2145 - 0750 


50.00 




12.50 


J1022 + 10 


23.00 




8.28 


50655 + 64 


5.00 




1.15 



Table 4: Densities in g cm ^for the hydrogen envelope of a 4 Mq star of radius IOORq. From 
Holmer (1998). 



r/RQ P (Poo)er 

1 

0.95 5.0(-ll) 7.8(-13) 

0.90 3.5(-9) 8.5(-13) 

0.85 5.8(-8) 9.3(-13) 

0.80 3.9(-7) 10.(-13) 



-25- 



I ' ^ 
PSR B1518+49 

PSR B1518+49 companion 

PSR B1534+12 

PSR B1534+12 companion 

PSR B1913+16 

PSR B1913+16 companion 

PSR B2127+11C 

PSR B2127+11C companion 

PSR B2303+46 

PSR B2303+46 companion 



PSR J0437-4715 
PSR J1012+5307 
PSR J 1045-4509 
PSR J1713+07 
PSR B1802-07 
PSR J1804-2718 
PSR B1855+09 
PSR J2019+2425 



lEil 



J_ 



J L 



I 
I 

I • 

I 

t 



PSR J0045-7319 

I I I L. 







1 2 
Neutron star mass (Mq) 



Fig. 1. — Neutron star masses from observations of radio pulsar system (Thorsett & 
Chakrabarty 1999). All error bars indicate central 68% confidence limits, except upper 
limits are one-sided 95% confidence limits. Five double neutron star systems are shown at 
the top of the diagram. In two cases, the average neutron star mass in a system is known 
with much better accuracy than the individual masses; these average masses are indicated 
with open circles. Eight neutron star-white dwarf binaries are shown in the center of the 
diagram, and one neutron star-main sequence star binary is shown at bottom. Vertical lines 
are drawn at m = 1.35 ± 0.04 Mq. As noted in our paper, PSR B2303-I-46 has since been 
shown to have a white dwarf companion.